EMIS/ELibM Electronic Journals

Outdated Archival Version

These pages are not updated anymore. They reflect the state of 20 August 2005. For the current production of this journal, please refer to http://www.math.psu.edu/era/.



% **************************************************************************
% * The TeX source for AMS journal articles is the publisher's TeX code    *
% * which may contain special commands defined for the AMS production      *
% * environment.  Therefore, it may not be possible to process these files *
% * through TeX without errors.  To display a typeset version of a journal *
% * article easily, we suggest that you either view the HTML version or    *
% * retrieve the article in DVI, PostScript, or PDF format.                *
% **************************************************************************
% Author Package file for use with AMS-LaTeX 1.2
% Date: 28-MAY-1996
%   \pagebreak: 0   \newpage: 0   \displaybreak: 0
%   \eject: 0   \break: 0   \allowbreak: 0
%   \clearpage: 0   \allowdisplaybreaks: 0
%   $$: 0   {eqnarray}: 2
%   \linebreak: 0   \newline: 0
%   \mag: 0   \mbox: 1   \input: 0
%   \baselineskip: 0   \rm: 0   \bf: 9   \it: 0
%   \hsize: 0   \vsize: 0
%   \hoffset: 0   \voffset: 0
%   \parindent: 0   \parskip: 0
%   \vfil: 0   \vfill: 0   \vskip: 0
%   \smallskip: 0   \medskip: 0   \bigskip: 0
%   \sl: 0   \def: 1   \let: 0   \renewcommand: 2
%   \tolerance: 0   \pretolerance: 0
%   \font: 0   \noindent: 0
%   ASCII 13 (Control-M Carriage return): 0
%   ASCII 10 (Control-J Linefeed): 0
%   ASCII 12 (Control-L Formfeed): 0
%   ASCII 0 (Control-@): 0
% Special characters: 0
%
%
%\controldates{16-JUL-1996,16-JUL-1996,16-JUL-1996,23-JUL-1996}
\documentclass{era-l}
%\issueinfo{2}{1}{}{1996}

%\NeedsTeXFormat{LaTeX2e}
%     \theorembodyfont{\slshape}
        \newtheorem{thm}{Theorem}[section]
     \newtheorem{prop}[thm]{Proposition}
     \newtheorem{lem}[thm]{Lemma}
     \newtheorem{cor}[thm]{Corollary}

 %    \theorembodyfont{\upshape}
\theoremstyle{definition} 
    \newtheorem{defn}[thm]{Definition}
     \newtheorem{notation}[thm]{Notation}
     \newtheorem{example}[thm]{Example}
     \newtheorem{conj}[thm]{Conjecture}
     \newtheorem{prob}[thm]{Problem}

\theoremstyle{remark}
     \newtheorem{rem}[thm]{Remark}

%\providecommand{\AMSMSC}[2]{
%\subjclass{Primary #1; Secondary #2.}}
\newcommand{\person}[1]{\textsc{#1}}

\hyphenation{di-men-sio-nal}
\newcommand{\comment}[1]{}
%\def
\newcommand\smnegskip{\mskip-0.6\thinmuskip}
\usepackage{amsfonts}
\newlength{\curmathsize}
%
%   Objects
%
\newcommand{\algebra}[1]{{{\mathfrak #1}}}
%% Change \Cliff, \object to simpler definitions since latex2html
%% doesn't yet handle optional arg newcommands properly [mjd,1996/07/30]
%%\newcommand{\Cliff}[2][\comment]{{\mathbf{Cl}(#1,#2)}}
\newcommand{\Cliff}[1]{{\mathbf{Cl}(#1)}}
%%\newcommand{\object}[2][\,]{{\mathrm{#2}#1}}
\newcommand{\object}[2]{{#1\mathrm{#2}}}
\newcommand{\Space}[2]{{ {{\mathbb #1}^{#2}} }}
\newcommand{\such}{\,\mid\,}
\newcommand{\FSpace}[2]{{{ #1_{#2} }}}
\newcommand{\bos}{{\mathcal B}}
%
%  Unary operations
%
\newcommand{\Cstar}{$C^{*}$}
\newcommand{\norm}[1]{\left\| #1 \right\|}
\newcommand{\modulus}[1]{\left|#1\right|}
%
%  Binary operations
%
\newcommand{\scalar}[2]{\langle #1,#2\rangle}
%
%\providecommand{\eqref}[1]{\textup{(\ref{#1})}}

\newcommand{\unit}{\mathbf{1}}
\renewcommand{\sp}[1]{\mathrm{sp}_{#1}\,}
\newcommand{\res}[1]{\mathrm{res}_{#1}\,}
\newcommand{\matr}[4]{{{ \left( \begin{array}{cc}
#1 & #2 \\ #3 & #4
\end{array}\right) }}}
%% Not used:
%%\newcommand{\spt}{\object{supp}}
\newcommand{\vecbf}[1]{\mathbf{#1}}
\newcommand{\algbf}[1]{ %\mbox
\text{\boldmath$\mathfrak{A}$}}

\begin{document}
\title[M\"obius transformations and monogenic functional calculus]
{M\"obius transformations and \\ monogenic functional calculus}
\author{Vladimir V. Kisil}
\address{Institute of Mathematics, %\\
                        Economics and Mechanics, %\\
                        Odessa State University, %\\
                        ul. Petra Velikogo, 2, %\\
                        Odessa-57, 270057, Ukraine}
\email{vk@imem.odessa.ua} %root@qchem.odessa.ua}

\copyrightinfo{1996}{American Mathematical Society}
\date{October 6, 1995, and, in revised form, March 9, 1996}
%\revdate{March 9, 1996}
\commby{Alexandre Kirillov}

\thanks{This work was partially supported by the INTAS grant 93-0322.
It was finished while the author enjoyed the hospitality of
Universiteit Gent, Vakgroep Wiskundige Analyse, Belgium.}
\keywords{Functional calculus, joint spectrum,  group representation,
intertwining operator, Clifford analysis, quantization} %.}
\subjclass{Primary 46H30, 47A13; Secondary 
30G35, 47A10, 47A60, 47B15, 81Q10}
\begin{abstract}
A new way of doing functional calculi is presented. A functional
calculus $\Phi: f(x)\rightarrow f(T)$ is not an algebra homomorphism
of a functional algebra into an operator algebra, but an intertwining
operator between two representations of a group acting on the two
algebras (as linear spaces).

This scheme is shown on the newly developed monogenic
functional calculus for an arbitrary set of non-commuting
self-adjoint operators. The corresponding spectrum and
spectral mapping theorem are included.
\end{abstract}
\maketitle
%\tableofcontents
\section{Introduction}

Functional calculus and the corresponding spectral theory belong to the
heart of functional
analysis~\cite{Helemskii93,SimonReed80,Vasilescu82}.
\person{J.~L.~Taylor}~\cite{JTaylor72} formulated a problem of functional
calculus in the following terms:

\textsl{For a given element $a$ of a Banach algebra $\algebra{A}$ the
correspondence $x\mapsto a$ gives rise to a representation of the
algebra of polynomials in the variable $x$ in $\algebra{A}$. It is
necessary to extend the representation to a larger functional
algebra\comment{ in the variable $x$}.}

The scheme works perfectly for several commuting variables $x_i$
and commuting elements $a_i\in\algebra{A}$, and gives rise to the
analytic Taylor calculus~\cite{JTaylor70,JTaylor72}.
But for the non-commuting case
the situation is not so simple: searching for an appropriate substitution for
a functional algebra in commuting variables was done in classes of Lie
algebras~\cite{JTaylor72}, freely generated algebras~\cite{JTaylor73},
cumbersome $\times$-algebras~\cite{KisRam95a}, etc. For the Weyl
calculus~\cite{Anderson69} no algebra homomorphism is known.

The main point of our approach (firstly presented in the paper)
is the following reformulation of the
problem, where the shift from an algebra \emph{homomorphism} to group
\emph{representations} is made:
\begin{defn}\label{de:calc}
Let a group algebra $\FSpace{L}{1}(G)$ have linear representations
${\pi}_{F}(g)$ in a space of functions $F$ and ${\pi}_B(g)$ in a Banach
algebra $B$ (depending on a set of elements $T\subset B$). One says
that a linear mapping $\Phi: F\rightarrow B$ is a {\em functional
calculus\/} if
$\Phi$ is an intertwining operator between ${\pi}_{F}$ and ${\pi}_B$,
namely
\begin{displaymath}
\Phi [{\pi}_{F}(g) f]={\pi}_B(g) \Phi[f],
\end{displaymath}
for all $g\in G$ and $f\in F$. Additional conditions will be stated later.
\end{defn}
One may ask whether there is a connection between these two
formulations at all. It is possible to check that the classic
Dunford-Riesz calculus (see~\cite[Chap.~IX]{RieszNagy55}
and~\cite[Chap.~2]{Helemskii93}) is generated
by the group of fractional-linear transformations of the complex line;
the Weyl~\cite{Anderson69} and continuous functional
calculi~\cite[Chap.~VII]{SimonReed80} are generated by the affine
group of  the real line; the analytic functional
calculus~\cite{JTaylor70,Vasilescu82} is based on
biholomorphic automorphisms of a
domain $U\subset\Space{C}{n}$;
and the monogenic (Clifford-Riesz) calculus~\cite{KisRam95a}
is generated by the group of
M\"obius transformations in $\Space{R}{n}$. Moreover, both the Weyl and
the analytic calculi can be obtained from the monogenic one by
restrictions of the symmetry group (see Theorem~\ref{th:space} and
Corollary~\ref{co:holom}). Such connections are based
on the role of group representations in function theories
(particularly on links with integral
representations~\cite{Kisil95d,Kisil95a}).

We will postpone the general results~\cite{Kisil95e} about functional
calculi. The subject of this paper is the construction of the
monogenic calculus from the M\"obius transformations (for other calculi see
Remark~\ref{re:other}). The monogenic and the
Riesz-Clifford~\cite{KisRam95a} calculi are cousins but not twins:
besides having conceptually different starting points,
they are technically different too. For example, for a
function $\Space{R}{n}\rightarrow \Cliff{n}$ we construct a calculus for
$n$-tuple of operators, not for $(n-1)$-tuples as it was done
in~\cite{KisRam95a}.

Due to the brief nature of present paper, ``proofs'' mean ``hints'' or
``sketches of proof''.

\section{Clifford analysis and the conformal group}
\comment{
In this Section we give notations and preliminary results. }

\subsection{The Clifford algebra and the M\"obius group}
Let $\Space{R}{n}$ be a Euclidean $n$-dimensional vector space with
a fixed frame $a_1$, $a_2$, \ldots, $a_n$ and let $\Cliff{n}$ be
the \emph{real Clifford algebra}~\cite[\S~12.1]{MTaylor86} generated by
$1$, $e_j$, $1\leq j\leq n$ and the relations
\begin{displaymath}
e_i e_j + e_j e_i =-2\delta_{ij}, \qquad 1e_i=e_i 1=e_i.
\end{displaymath}
We put $e_0=1$. The embedding $\algebra{i}: \Space{R}{n}\rightarrow
\Cliff{n}$ is defined by the formula:
\begin{equation}\label{eq:embedding}
\algebra{i}: x=\sum_{j=1}^n x_j a_j \mapsto \vecbf{x}=\sum_{j=1}^n x_j e_j.
\end{equation}
We identify $\Space{R}{n}$ with its image under $\algebra{i}$ and
call its elements \emph{vectors}. There are two linear
anti{-}automorphisms $*$ and $-$ of $\Cliff{n}$ defined on its basis
$A_\nu=e_{j_1}e_{j_2}\cdots e_{j_r}$, $1\leq j_1 <\cdots2$ is
not so rich as the conformal group in $\Space{R}{2}$.
Nevertheless, the conformal covariance has many applications in
Clifford analysis~\cite{Ryan95b}.
Notably,
groups of conformal mappings of open unit balls $\Space{B}{n} \subset
\Space{R}{n}$ \emph{on}to itself are similar for all $n$ and
as sets can be
parametrized by the product of $\Space{B}{n}$ itself and the group
of isometries of its boundary $\Space{S}{n-1}$.
\begin{lem}
Let $a\in\Space{B}{n}$, $b\in\Gamma_n$; then the M\"obius transformations of
the form
\begin{displaymath}
\phi_{(a,b)}=\matr{b}{0}{0}{{b}^{*-1}}\matr{1}{-
a}{{a}^*}{-1}=\matr{b}{-ba}{{b}^{*-1}a^*}
{-{b}^{*-1}}
\end{displaymath}
constitute the group $B_{n}$ of conformal mappings of the open unit ball
$\Space{B}{n}$ onto itself. $B_{n}$ acts on
$\Space{B}{n}$ transitively.
Transformations of the form $\phi_{(0,b)}$  constitute a
subgroup isomorphic to $\object{O}{(n)}$. The homogeneous space
$B_{n}/\object{O}{(n)}$ is isomorphic as a set to
$\Space{B}{n}$. Moreover:
\begin{enumerate}
\item $\phi_{(a,1)}^2=1$ identically on $\Space{B}{n}$
($\phi_{(a,1)}^{-1}=\phi_{(a,1)}$).
\item $\phi_{(a,1)}(0)=a$, $\phi_{(a,1)}(a)=0$.
\comment{
\item $\phi_{(a,1)}\phi_{(b,1)}=\phi_{(c,d)}$ where
$c=\phi_{(b,1)}(a)$ and $=\phi_{(a,1)}$ }
\end{enumerate}
\end{lem}

Obviously, conformal mappings preserve the space of harmonic
functions, i.e., null solutions to the \emph{Laplace} operator
$\Delta=\sum_{j=1}^n \frac{\partial^2 }{\partial x_j^2}$. They also
preserve the space of \emph{monogenic} functions, i.e.,
null solutions $f: \Space{R}{n}\rightarrow
\Cliff{n}$ of the \emph{Dirac}~\cite{BraDelSom82,DelSomSou92} operator $D=\sum_{j=1}^n e_i
\frac{\partial }{\partial x_j}$ (note that $\Delta=D\bar{D}=-D^2$). The
group $B_{n}$ is sufficient for construction of the Poisson
integral representation of harmonic functions and the Cauchy and
Bergman formulas in Clifford analysis by the formula~\cite{Kisil95d}
\begin{equation}\label{eq:reproduce}
K(x,y)=c\int_G [\pi_g f](x) \overline{[\pi_g f](y)}\,dg,
\end{equation}
where $\pi_g$ is an irreducible unitary
square integrable representation~\cite[\S~9.3]{Kirillov74} of a group $G$,
$f(x)$ is an arbitrary non-zero function, and $c$ is a constant.

\section{Monogenic calculus}
\subsection{Representations of the M\"obius group in \Cstar-algebras}
Fix an $n$-tuple of \emph{bounded self-adjoint} elements
$T=(T_1,\ldots,T_n)$ of a \Cstar-algebra $\algebra{A}$.  Let
${\algbf{A}}=\algebra{A}\otimes \Cliff{n}$. We can associate with $T$
an element $\vecbf{T}\in {\algbf{A}}$ by the
formula~\cite{McInPryde87}
\begin{displaymath}
\vecbf{T}=\sum_{j=1}^n T_j\otimes e_j
\end{displaymath}
in analogy with the embedding $\algebra{i}$ of~\eqref{eq:embedding}.
\comment{
We again call elements of $\algbf{A}$ in the above form by
\emph{vectors}.
We combine two natural anti{-}automorphisms $*$ and $-$ on
$\algebra{A}$ and $\Cliff{n}$ and define a linear anti-automorphism
on  ${\algbf{A}}$ by its action on elementary
tensors:
\begin{displaymath}
\overline{A\otimes a}=A^* \otimes \bar{a}, \qquad A\in \algebra{A}, \
a\in\Cliff{n}.
\end{displaymath}
Particularly  $\vecbf{T}\widetilde{\vecbf{T}}=
\widetilde{\vecbf{T}}\vecbf{T}= \sum_{j=1}^n \norm{T_j}^2$ is real we
would denote it by $\norm{\vecbf{T}}^2$. }
One notes the naturality of the following
\begin{defn}
The \emph{Clifford resolvent set} $R(\vecbf{T})$ of an $n$-tuple
$T$ is the maximal open subset of $\Space{R}{n}$ such that for
$\lambda\in R(\vecbf{T})$ the element $\vecbf{T}-\lambda I$ is invertible in
${\algbf{A}}$.
The \emph{Clifford spectrum} $\sigma(\vecbf{T})$ is the completion of the
Clifford resolvent set $\Space{R}{n}\setminus R(\vecbf{T})$.
\end{defn}
For example, the joint spectrums of one $\{j_1\}$, two $\{j_1,j_2\}$,
three $\{j_1,j_2,j_3\}$ Pauli matrices~\cite[\S~12.4]{MTaylor86} are
the unit sphere $\Space{S}{0}$ ($=\{-1,1\}$), the origin $(0,0)$ (= a sphere
of zero radius), the unit sphere $\Space{S}{2}$ in $\Space{R}{1}$,
$\Space{R}{2}$, and $\Space{R}{3}$, respectively.
The canonical connection between the Grassmann and Clifford algebras
gives (see~\cite[\S~III.6]{Vasilescu82})
\begin{lem}
Let an $n$-tuple $T$ consist of mutually commuting operators in
$\algebra{A}$. Then the Taylor joint spectrum\footnote{There are less
used non-commutative versions of the Taylor joint
spectrum~\cite{JTaylor72}.}~\cite{JTaylor70} and the Clifford spectrum
coincide.
\end{lem}
Let $V$ be a subgroup of the Vahlen group $V(n)$ such that for any
$\matr{a}{b}{c}{d}\in V$ we have $c^{-1}d \in R(\vecbf{T})$ (i.e.,
$c\vecbf{T}+d$ is invertible in $\algbf{A}$). Then there is a
(non-linear) representation $\pi_T$ of $V$:
\begin{equation}\label{eq:op-rep}
\pi_{T}\matr{a}{b}{c}{d}: \vecbf{T} \mapsto
(a\vecbf{T}+b)(c\vecbf{T}+d)^{-1}.
\end{equation}
The following easy lemma has a useful corollary.
\begin{lem}\label{le:spectr}
A vector $\lambda$ belongs to the resolvent set $R(\vecbf{T})$ if and only
if $\phi_{(\lambda,1)}(\vecbf{T})$ is well defined.
\end{lem}
\begin{cor}\label{co:spectr}
A vector $\lambda$ belongs to the spectrum $\sigma(\vecbf{T})$ if and only
if the vector $\phi_{(\lambda,1)}(\beta)$ belongs to the spectrum
$\sigma(\phi_{(\beta,1)}(\vecbf{T}))$.
\end{cor}
\comment{
\begin{defn}
An intertwining operator $\Theta$ between two representations
$\pi_{\Space{R}{n}}$~\eqref{eq:sp-rep} and $\pi_{\algebra{A}}$
is called a \emph{coordinate mapping}.
\end{defn}
}
Having a representation $\pi$ of the group $G$ in a \emph{linear} space
$X$ one can define the \emph{linear} representation $\pi_L$ of the
convolution algebra $\FSpace{L}{1}(G)$ in the space of mappings
$X\rightarrow X$ by the rule
\begin{equation}\label{eq:extension}
\pi_L(f)= \int_G f(g) \pi(g)\, dg,
\end{equation}
where $dg$ is Haar measure on $G$.

We denote such an extension of
$\pi_{\Space{R}{n}}$ of~\eqref{eq:sp-rep} by $\pi_F$ and the extension
of $\pi_{T}$ of~\eqref{eq:op-rep} by $\pi_{\algebra{A}}$.
Let $\FSpace{F}{}(\Space{R}{n})=\pi_F(\FSpace{L}{1}(G))$,
$\widehat{f}(\vecbf{x})=\pi_F(f)$,
$\widehat{f}(\vecbf{T})=\pi_\algebra{A}(f)$.
Following Definition~\ref{de:calc} we give
\begin{defn}
A \emph{monogenic functional calculus}
$\Phi_V:\FSpace{F}{}(\Space{R}{n})\rightarrow \algbf{A}$ for
a subgroup $V$ and $n$-tuple $T$ is a linear operator with the
following properties:
\begin{enumerate}
\item\label{it:fc-rep} It is an intertwining operator between the two
representations
$\pi_F$ and $\pi_{\algebra{A}}$:
\begin{displaymath}
\Phi_V \pi_{F} =\pi_{\algebra{A}} \Phi_V.
\end{displaymath}
\item\label{it:fc-iden} $\Phi \widehat{\unit}=I$, i.e., the image of the
function identically equal to $1$ is the unit operator.
\item\label{it:fc-t} $\Phi \widehat{\delta}_e=\vecbf{T}$, where $\delta_e$ is
the delta function centered at the group unit.
\end{enumerate}
\end{defn}
The following theorem is group{-}independent:
\begin{thm}\label{th:compose}
Let $k(g)$ and $l(g)$ be functions on $V$ such that
both $\widehat{l}(\vecbf{T})$ and
$\widehat{k}(\widehat{l}(\vecbf{T}))$ are well defined.
Then $\widehat{k}(\widehat{l}(\vecbf{T}))=
[\widehat{k*l}](\vecbf{T})$ ($*$ is convolution on the
group). In particular, $\pi_\algebra{A}(g)[\widehat{l}(\vecbf{T})]=
[\widehat{\pi_l(g)l}](\vecbf{T})$.
\end{thm}
\begin{thm}[Uniqueness]
If the representation $\pi_F$ is irreducible, then
the functional calculus (if any) is unique.
Particularly, for a given simply connected domain $\Omega$ and
an $n$-tuple of operators $T$, there exists no more than one
monogenic calculus.
\end{thm}
\begin{thm}\label{th:space}
For any $n$-tuple $T$ of bounded self-adjoint operators there
exists a monogenic calculus on $\Space{R}{n}$. This calculus coincides on
monogenic functions
with the Weyl functional calculus from~\cite{Anderson69}.
Particularly, the polynomial functions of operators are the symmetric
(Weyl) polynomials.
\end{thm}
\begin{proof}
M\"obius transformations of $\Space{R}{n}$ (without the point $\infty$)
are the affine transformations of $\Space{R}{n}$, which are
represented via
upper-triangular Vahlen matrices. But the Weyl calculus is defined
uniquely by its affine covariance~\cite[Theorem~2.4(a)]{Anderson69}
according to the following lemma, which is given without proof.
\end{proof}
\begin{lem}
The one-dimensional Weyl calculus is uniquely defined by its affine
covariance.
\end{lem}

The following result is a direct counterpart of the classic one.
\begin{thm}
For the existence of the monogenic functional calculus in the unit ball
$\Space{B}{n}$ it is necessary that $\sigma(\vecbf{T})\subset
\Space{B}{n}$.
\end{thm}
In the next subsection we obtain via an integral representation that this
condition is also sufficient.

\subsection{Integral representation}

The best way to obtain an integral formula for the monogenic calculus
is to use~\eqref{eq:reproduce}. To save space we will now
use  another straightforward approach, which however could be
obtained from~\eqref{eq:reproduce} by a suitable representation of
$\Space{Z}{}$.
We already know from Theorem~\ref{th:space} how to construct monogenic
functions for polynomials. Thus, by the linearity (via
decomposition~\cite[Chap.~II, (1.16)]{DelSomSou92}) there is only one
way to define the Cauchy kernel of the operators $T_j$.
\begin{defn} Let us define
(we use notation
from~\cite{DelSomSou92})\begin{equation}\label{eq:cauchy-op}
E(y,T)=\sum_{j=0}^\infty\left(\sum_{\modulus{\alpha }=j}V_\alpha
(T)W_\alpha (y)\right) ,
\end{equation}
where
\begin{align} %\begin{eqnarray}
W_\alpha (x)&=(-1)^{\modulus{\alpha}}\partial ^\alpha E(x)
=\frac{(-1)^{\modulus{\alpha}}\Gamma(\frac{n+1}{2})}{2\pi^{(n+1)/2}}\,
\partial ^\alpha \frac{\overline{x}}{\modulus{x}^{n+1}},\nonumber\\
V_\alpha(T) &= \frac{1}{\alpha !}\sum_\sigma (T_1 e_{\sigma(1)}-T_{\sigma(1)})
(T_1e_{\sigma(2)}-T_{\sigma(2)})\cdots (T_1e_{\sigma(n)}- T_{\sigma(n)}),
\label{eq:sym-pol}
\end{align} %\end{eqnarray}
where the summation is taken over all possible permutations.
\end{defn}
\begin{lem}\label{le:bound}
Let $\modulus{T}=\lim_{j\rightarrow\infty}\sup_\sigma
\norm{T_{\sigma(1)}
\cdots T_{\sigma(j)} }^{1/j}$, $1\leq\sigma(i)\leq n$ be the
Rota-Strang joint
spectral radius~\cite{Rota60a}.
Then for a fixed $\modulus{y}>\modulus{T}$,
equation~\eqref{eq:cauchy-op}
defines a bounded operator in $\algbf{A}$.
\end{lem}
As usual, by Liouville's theorem~\cite[Theorem~5.5]{McInPryde87}
for the function $E(y,T)$ the Clifford
spectrum is not empty; by definition it is closed and by the previous
lemma it is bounded. Thus
\begin{lem} The Clifford spectrum $\sigma(\vecbf{T})$ is compact.
\end{lem}
\begin{lem}\label{le:v-lem}
Let $r>\modulus{T}$ and let $\Omega$ be the ball $\Space{B}{}(0,r)\in
\Space{R}{n}$. Then \comment{for any symmetric polynomial $P(\vecbf{x})$} we
have
\begin{equation}\label{eq:int-polyn}
P(T)=\int_{\partial \Space{B}{}} E(y,T)\, dn_y\, P(y),
\end{equation}
where $P(T)$ is the symmetric polynomial~\eqref{eq:sym-pol} of the
$n$-tuple $T$.
\end{lem}
\begin{lem}
For any domain $\Omega$ that does not contain $\sigma(\vecbf{T})$ and any
$f\in\FSpace{H}{}(\Omega)$, we have
\begin{equation}
\int_{\partial \Omega} E(y,T)\, dn_y\, f(y)=0.
\end{equation}
\end{lem}

Due to this lemma we can replace the domain  $\Space{B}{}(0,r) $ in
Lemma~\ref{le:v-lem} with an arbitrary domain $\Omega$
containing the spectrum $\sigma(\vecbf{T}).$
An application of Lemma~\ref{le:v-lem} gives the
main
\begin{thm}
Let $T=(T_1,\ldots,T_n)$ be an $n$-tuple of bounded self-adjoint
operators. Let the domain $\Omega$ with piecewise smooth boundary have
a connected complement and suppose the spectrum $\sigma(\vecbf{T})$ lies
inside the domain $\Omega$. Then the mapping
\begin{equation}\label{eq:int-calc}
\Phi: f(x)\mapsto f(T)=\int_{\partial \Omega} E(y,T)\, dn_y\,
f(y)
\end{equation}
defines the monogenic calculus for $T$.
\end{thm}
Because holomorphic functions are also monogenic,
we can apply our construction to them. For a commuting
$n$-tuple $T$ the calculus constructed above will coincide with
the analytic one~\cite{Vasilescu82} on the polynomials. This gives
\begin{cor}\label{co:holom}
Restriction of the monogenic calculus for a commuting
$n$-tuple $T$ to holomorphic functions produces
the analytic functional calculus.
\end{cor}

\subsection{Spectral mapping theorem}
A good notion of functional calculus should be connected with a good
notion of spectrum via the spectral mapping theorem%
\comment{ and spectral decomposition}.
In the analytic calculus~\cite[II.2.23]{Helemskii93} and
the functional calculus of self-adjoint
operators~\cite[Theorem VII.1(e)]{SimonReed80} the homomorphism property
plays a crucial role in proofs of spectral mapping theorems. Nevertheless,
in our versions these results are also preserved.

It is easy to check that formula~\cite[(3.5)]{Ahlfors86}
can be adapted in the following way:
\begin{lem}
Let $(-c^{-1}d)\not\in \sigma(\vecbf{T})$ and $\vecbf{x}\not=-c^{-
1}d$.
Let $g$ be the action of $\matr{a}{b}{c}{d}$ in $\algbf{A}$. Then
\begin{displaymath}
g(\vecbf{T})-g(\vecbf{x})=(c\vecbf{x}+d)^{*-1}(\vecbf{T}-
\vecbf{x})(c\vecbf{T}+d)^{-1}.
\end{displaymath}
\end{lem}
\begin{thm}[Spectral mapping theorem]
\begin{displaymath}
\sigma(\widehat{f}(\vecbf{T}))=\{\widehat{f}(\vecbf{x}) \such
\vecbf{x}\in\sigma(\vecbf{T})\}.
\end{displaymath}
\end{thm}
\begin{proof}
Let $x\in\sigma(\vecbf{T})$ and $\vecbf{y}=f(\vecbf{x})$ then
\begin{align*} %\begin{eqnarray*}
\widehat{f}(\vecbf{T})-\vecbf{y}&=\widehat{f}(\vecbf{T})-\widehat{f}
(\vecbf{x})\\
  &=\int_V f(g) g(\vecbf{T})\, dg -\int_V f(g) g(\vecbf{x})\, dg\\
  &=\int_V f(g) \left[g(\vecbf{T})-g(\vecbf{x})\right]\, dg \\
  &=\int_V f(g) (c\vecbf{x}+d)^{*-1}(\vecbf{T}-
\vecbf{x})(c\vecbf{T}+d)^{-1}\, dg.
\end{align*} %\end{eqnarray*}
Thus $f(\vecbf{T})-\vecbf{y}$ belongs to the closed proper left ideal
generated by $\vecbf{T}-\vecbf{x}$.

Otherwise let $\widehat{f}(\vecbf{x})\not = \vecbf{y}$ for all $\vecbf{x}
\in \Omega \supset \sigma(\vecbf{T})$. Analogously to Lemma~\ref{le:spectr}
with the help of Theorem~\ref{th:compose} we conclude that the function
\begin{displaymath}
\widehat{[\phi_{(y,1)}f]}(\vecbf{x})=\phi_{(y,1)}\int_V f(g) g(\vecbf{x})\, dg=\int_V f(\phi_{(y,1)}g) g(\vecbf{x})\, dg
\end{displaymath}
is well defined on $\Omega$. Thus we can define an operator
\begin{displaymath}
\widehat{[\phi_{(y,1)}f]}(\vecbf{T})=\int_V f(\phi_{(y,1)}g) g(\vecbf{T})\, dg.
\end{displaymath}
Due to Lemma~\ref{le:spectr} the existence of such an operator
provides us with the invertibility of the operator $\widehat{f}(\vecbf{T}) -\vecbf{y}$.
\end{proof}

\section{Concluding remarks}
\begin{rem}
\textup{It seems worthwhile (if not very natural) to connect the two main
motives
of functional analysis: functional calculus and group representations.
Such a connection causes changes in applications based on the
functional
calculi (quantum mechanics~\cite{Kisil95g}): to quantize one could use not
algebra homomorphisms (observables are usually believed to form an
algebra) but representations of groups (symmetry of the system
under consideration).}
\end{rem}
\begin{rem}
\textup{There is the Cayley transformation~\cite{Ahlfors86} of
the unit ball to the
``upper half-plane'' $x_{n}\geq 0$. This allows one to make a
straightforward modification of the monogenic calculus for
an $n$-tuple of operators where only $T_{n}$ is
semibounded (Hamiltonian!) and the other operators are just unbounded.}
\end{rem}
\begin{rem}\label{re:other}
\textup{The classic
Dunford-Riesz~\cite[Chap.~IX]{RieszNagy55} calculus can be obtained
similarly using the representation of
$\object{SL}{(2,\Space{R}{})}$ by fractional-linear mappings~\cite{MTaylor86}
of the complex
line $\Space{C}{}$. The functional calculus of a self-adjoint
operator~\cite[Chap.~VII]{SimonReed80} can be obtained from
affine transformations of the real line. But a non-formal integral formula
is possible only through an integral representation of the $\delta$-function:
in this way we have arrived, for example, at the Weyl functional
calculus~\cite{Anderson69}.}
\textup{Biholomorphic automorphisms of unit ball in $\Space{C}{n}$
also form a sufficiently large group (subgroup of the M\"obius group) to
construct a function theory~\cite{Kisil95a}. Thus, we can develop a
functional calculus based on the theory of several complex variables.
Due to its biholomorphic
covariance~\cite{CurtoVasil94a,CurtoVasil93a} it must coincide
with the well-known analytic calculus~\cite{Vasilescu82}.}
\end{rem}
\begin{rem}
\textup{Our Definition~\ref{de:calc} is connected with
\emph{Arveson-Connes spectral
theory}~\cite{\comment{Arveson80,Connes73,}Takesaki83}. The main differences
are: (i) most of our groups are non-commutative;
(ii) we use a space of analytic functions (not functions on the group
directly) as a model of functional calculus and definition of a
spectrum.
It seems that the first feature is mainly due to the second
one.}
\end{rem}

\comment{Our technique is mostly
elementary unlike the finest homological approach of J.~L.~Taylor.}
\newcommand{\noopsort}[1]{} \newcommand{\printfirst}[2]{#1}
  \newcommand{\singleletter}[1]{#1} \newcommand{\switchargs}[2]{#2#1}
  \newcommand{\irm}{\textup{I}} \newcommand{\iirm}{\textup{II}}
  \newcommand{\vrm}{\textup{V}}
\makeatletter \renewcommand{\@biblabel}[1]{\hfill#1.}\makeatother
\begin{thebibliography}{10}

\bibitem{Ahlfors86}
L.~V. Ahlfors, {\em {M{\"o}bius} transformations in {$\Space{R}{n}$} expressed
  through $2\times 2$ matrices of {Clifford} numbers}, Complex Variables Theory
  Appl. %{\bf
\textbf{ 5} (1986), no.~2, 215--224.
 \MR{88a:15052} 
\bibitem{Anderson69}
R.~F.~V. Anderson, {\em The {Weyl} functional calculus}, J. Funct. Anal. %{\bf
\textbf{
  4} (1969), 240--267.
 \MR{58:30405} 
\comment{\bibitem{Arveson80}
W.~Arveson, {\em The harmonic analysis of automorphisms groups}, American
  Mathematical Society, Amer. Math. Soc. Summer Institute,
{\noopsort{}}1980.}

\bibitem{BraDelSom82}
F.~Brackx, R.~Delanghe, and F.~Sommen, {\em Clifford analysis}, Research Notes
  in Mathematics, vol.~76, Pitman Advanced Publishing Program, Boston, 1982.
 \MR{85j:30103}
\bibitem{Cnops94a}
J.~Cnops, {\em {Hurwitz} pairs and applications of {M\"obius} transformations},
  {Habilitation} dissertation, Universiteit Gent, Faculteit van de
  Wetenschappen, 1994.

\comment{\bibitem{Connes73}
A.~Connes, {\em Une classification des facteurs de type{ \textup{III}}}, Ann.
  Sci. {\'E}cole Norm. Sup. (4) %{\bf
\textbf{ 6} (1973), 133--252.}
\bibitem{CurtoVasil94a}
R.~E. Curto and F.-H. Vasilescu, {\em Standard operator models in several
  variables}, preprint.

\bibitem{CurtoVasil93a}
\bysame, {\em Automorphism invariance of the operator-valued {Poisson}
  transform}, Acta Sci. Math. (Szeged) %{\bf
\textbf{ 57} (1993), 65--78.
 \MR{94i:47013}
\bibitem{DelSomSou92}
R.~Delanghe, F.~Sommen, and V.~Sou\v{c}ek, {\em Clifford algebra and
  spinor-valued functions}, Kluwer Academic Publishers, Dordrecht,
  {\noopsort{}}1992.
 \MR{94d:30084}
\bibitem{Helemskii93}
A.~Y. Helemskii, {\em Banach and locally convex algebras}, Clarendon Press,
  Oxford, {\noopsort{}}1993.
\MR{94f:46001}
\bibitem{Kirillov74}
A.~A. Kirillov, {\em Elements of the theory of representations},
  Springer-Verlag, New York, 1976.
 \MR{54:447} 
\bibitem{Kisil95d}
V.~V. Kisil, {\em Construction of integral representations in spaces of
  analytical functions}, Dokl. Akad. Nauk SSSR, to appear.

\bibitem{Kisil95g}
\bysame, {\em Do we need that observables form an algebra?}, in preparation.

\bibitem{Kisil95a}
\bysame, {\em Integral representation and coherent states}, Bull. Soc. Math.
  Belg. S{\'e}r. A \textbf{ 2} (1995), 529--540. % to appear.
\CMP{96:08}

\bibitem{Kisil95e}
\bysame, {\em Spectrum of operator, functional calculi and group
  representations}, in preparation.

\bibitem{KisRam95a}
V.~V. Kisil and E.~Ram\'{\i}rez~de Arellano, {\em The {Riesz-Clifford}
  functional calculus for several non-commuting operators and quantum field
  theory}, Math. Methods Appl. Sci., to appear.

\bibitem{McInPryde87}
A.~McIntosh and A.~Pryde, {\em A functional calculus for several commuting
  operators}, Indiana Univ. Math.~J. %{\bf
\textbf{ 36} (1987), 421--439.
 \MR{88i:47007}
\bibitem{SimonReed80}
M.~Reed and B.~Simon, {\em Functional analysis}, Methods of Modern Mathematical
  Physics, vol.~1, Academic Press, Orlando, second ed., {\noopsort{}}1980.
 \MR{85e:46002} 
\bibitem{RieszNagy55}
F.~Riesz and B.~Sz-Nagy, {\em Functional analysis}, Ungar, New York,
  {\noopsort{}}1955.
\MR{17:175i}
\bibitem{Rota60a}
G.-C. Rota and W. G.~Strang, {\em A note on the joint spectral radius}, Nederl. Akad. Wetensch.
  Indag. Math. %{\bf
\textbf{ 22} (1960), 379--381.
\MR{26:5434} 
\bibitem{Ryan95b}
J.~Ryan, {\em Some application of conformal covariance in {Clifford} analysis},
  {Clifford} Algebras in Analysis and Related Topics (J.~Ryan,
  ed.), CRC Press, Boca Raton, {\noopsort{}}1996, pp.~128--155.
\CMP{96:10}
\bibitem{Takesaki83}
M.~Takesaki, {\em Structure of factors and automorphism groups}, Regional
  Conference Series in Mathematics, vol.~51, American Mathematical Society,
  Providence, Rhode Island, {\noopsort{}}1983.
 \MR{84k:46043}
\bibitem{JTaylor70}
J.~L. Taylor, {\em The analytic-functional calculus for several commuting
  operators}, Acta Math. %{\bf
\textbf{ 125} (1970), 1--38.
 \MR{42:6622} 
\bibitem{JTaylor72}
\bysame, {\em A general framework for a multioperator functional calculus},
  Adv. Math. %{\bf
\textbf{ 9} (1972), 183--252.
 \MR{48:6967}
\bibitem{JTaylor73}
\bysame, {\em Functions of several noncommuting variables}, Bulletin of the
  American Mathematical Society %{\bf
\textbf{ 79} (1973), no.~1, 1--34.
 \MR{47:3995} 
\bibitem{MTaylor86}
M.~E. Taylor, {\em Noncommutative harmonic analysis}, Math. Surv. and
  Monographs, vol.~22, American Mathematical Society, Providence, Rhode Island,
  {\noopsort{}}1986.
 \MR{88a:22021} 
\bibitem{Vasilescu82}
F.-H. Vasilescu, {\em Analytic functional calculus and spectral decomposition},
  Mathematics and Its Applications, vol.~1, D.~Reidel Publ. Comp., Dordrecht,
  Holland, {\noopsort{}}1982.
 \MR{85b:47016} 
\end{thebibliography}

\comment{
\bibliographystyle{amsshort}
\bibliography{mrabbrev,analyst} }
\end{document}